Articles
Dynamics in diffusive Leslie–Gower prey–predator model with weak diffusion
Dynamics in diffusive Leslie–Gower prey–predator model with weak diffusion
Nonlinear Analysis: Modelling and Control, vol. 27, no. 6, pp. 1168-1188, 2022
Vilniaus Universitetas

Received: 14 January 2022
Published: 01 November 2022
Abstract: This paper is concerned with the diffusive Leslie–Gower prey–predator model with weak diffusion. Assuming that the diffusion rates of prey and predator are sufficiently small and the natural growth rate of prey is much greater than that of predators, the diffusive Leslie–Gower prey– predator model is a singularly perturbed problem. Using travelling wave transformation, we firstly transform our problem into a multiscale slow-fast system with two small parameters. We prove the existence of heteroclinic orbit, canard explosion phenomenon and relaxation oscillation cycle for the slow-fast system by applying the geometric singular perturbation theory. Thus, we get the existence of travelling waves and periodic solutions of the original reaction–diffusion model. Furthermore, we also give some numerical examples to illustrate our theoretical results.
Keywords: Leslie–Gower prey–predator model, geometric singular perturbation theory, relaxation oscillation, canard explosion phenomenon, heteroclinic orbit.
1 Introduction
The abundant dynamical feature of interacting species is a hot issue in the research of ecological system. Based on laboratory experiments and observations, researchers have proposed many realistic biological models such as prey–predator model. Among these models, a typical one is the Leslie–Gower prey–predator model, which is firstly proposed in [19]. In the Leslie–Gower prey–predator model, the density of prey can influence the carrying capacity of predators such that the density of predators obeys a logistic
| Parameter | Interpretation | Unit |
| K | Carrying capacity of the prey | N |
| r1 | Natural growth rate of the prey | time−1 |
| r2 | Natural growth rate of the predator | time−1 |
| b1 | The maximum value of the per capita reduction rate of the prey | time−1 |
| b2 | The maximum value of the per capita reduction rate of the predator | time−1 |
| m1 | Environment protection to the prey | N |
| m2 | Environment protection to the predator | N |
dynamics with a changing capacity proportional. Collings [7] highlights that the Leslie– Gower prey–predator model is sufficient in population dynamics because it can avoid the biological control paradox of classical prey–predator models—a prey density is low in a stable coexistence equilibrium. The modified Leslie–Gower predator–prey model is proposed as
where U and V separately stand for the total numbers of preys and predators, all parameters are positive constants and their interpretations are described in Table 1.
The distributions of prey and predators are not homogeneous in real world. Thus, we introduce the diffusive terms into model (1), which describe the movement of preys and predators, and obtain the following modified Leslie–Gower reaction–diffusion model:
where and are diffusion rate of the prey and predators, respectively.
In recent years, many researchers have investigated the modified Leslie–Gower predator–prey model (1) and its reaction–diffusion case [1, 2, 14, 15, 25, 26, 29, 33]. By using the perturbation methods, they derived some interesting conclusions such as the stability of equilibriums, bifurcations, travelling waves, periodic solutions and so on.
For mathematical simplicity, using the following scaling transformations
and denoting and by and , we can get nondimensional model (2) as
We study model (3) under assumption that the diffusion rates of predators and prey are similar and sufficiently small and prey grow much faster than predators. It is clear that this assumption is reasonable such as hares and coyotes. In this article, we mainly investigate the travelling waves and periodic solutions of model (3). Hence, using the travelling wave transformation [9, 10], we get
where c stands for the velocity of travelling waves. Note that the solution of (4) with represents standing waves, which is not our main focus in this article. Since system (4) is invariant under transformation , then we only need to consider the case . Furthermore, in this paper, we are keen on the travelling waves with , which means the velocity c is larger than the week diffusion rate and . Hence, by transform , the equivalent system of (4) reads
Under our assumption, we introduce following new notations:
and we rewrite (5) as a singularly perturbed system
where and are sufficiently small parameters, and is called the slow variable. Furthermore, we assume that and satisfy .
For the fast variable , we get
Note that systems (6) and (7) are separately called the slow system and the fast system, and their dynamics are equivalent.
The slow-fast system is an important component in biological models and investigated by many investigators, who obtained some new dynamics. For prey–predator model, Ambrosio et al. [3] and Zhang and Wang [32] separately studied the relaxation oscillation cycle of system (1) with one small parameter. In [20], Li and Zhu investigated the canard cycles and relaxation oscillation cycle of a slow-fast prey–predator system with Holling III and IV response functions. Furthermore, in [28], Shen investigated the dynamics of similar model with Holling IV response function. Atabaigi and Barati [4] also investigate the relaxation oscillation cycle and canard explosion phenomenon of a slow-fast Holling– Tanner model with Holling IV response function. Indeed, Zhang and Wang [31] studied the dynamics of a slow-fast Holling–Tanner model with Holling III response function. For high dimension model, Liu, Xiao and Yi [23] considered the relaxation oscillation cycle of a slow-fast prey-predator model with one prey and two competing predators, and Shen, Hsu and Yang [27] studied the dynamics of a slow-fast intraguild predation model. For reaction–diffusion cases, Ducrot, Liu and Magal [11] studied the large speed traveling waves for the diffusive Rosenzweig–MacArthur predator–prey model, and Cai, Ghazaryan and Manukian [5] studied the travelling waves for the diffusive Rosenzweig– MacArthur and Holling–Tanner models with two small parameters. Furthermore, the geometric singular perturbation theory is an useful analysis method in these paper.
Recently, the geometric singular perturbation theory has been a main method in the analysis of a slow-fast system and it contains many mathematical tools such as the Fenichel theory [12, 18], the exchange lemma [21, 22], the blow-up method [16, 17] and the entryexit function [8,30] and so on. Whereas, the foundation of geometric singular perturbation theory is the Fenichel theory about locally invariant manifolds. Its main conclusion is that, when , there is a locally invariant slow manifold in-neighborhood of a normally hyperbolic submanifold of the critical manifold .
In this article, we will apply the methods and conclusions in geometric singular theory to analyse (6) and (7). We obtain the existence of heteroclinic orbits corresponding to travelling waves and canard cycles and relaxation oscillation cycle corresponding to periodic solutions of system (3). Furthermore, compared with the result for two-dimensional space in [3, 32], we analyse the canard cycles and relaxation oscillation cycle of the four-dimensional slow-fast system (6) with two small parameters. Indeed, we also illustrate the canard explosion phenomenon which is the changing process from a small limit cycle in the singular Hopf bifurcation to the relaxation oscillation cycle.
The rest of this article is organized as follows. In Section 2, we reduce system (6) to the plane . We analyse the existence of travelling waves in Section 3 and prove the existence of canard explosion phenomenon and relaxation oscillation cycle in Section 4. Finally, we give the conclusion in Section 5.
2 Equilibriums and critical manifold
Under our assumption , it is clear that system (6) is a multiscale slowfast system. Thus, using the method in [5, 13, 24], we reduce the multiscale slow-fast system (6) twice by the geometric singular perturbation. The first reduction is respect to smaller parameter , and the second is respect to .
For , the degenerate system of (6) reads
and we have the critical manifold
which is the set of equilibriums for the layer system of (7) as
It is easy to verify each point of has eigenvalues -1 and besides the two zero eigenvalues. Hence, the critical manifold is normally hyperbolic and attracting based on the Fenichel theory [12, 18]. Furthermore, the dynamics in critical manifold are defined by the limiting system
which is a slow-fast system respect to small parameter
Based on the geometric singular perturbation theory [6,18], there is a two dimensional stable slow manifold for , which is viewed as a .-order perturbation of the critical manifold , i.e.,. Furthermore, the flows on are govern by the -perturbation of system (8). Hence, we only need to investigate system (8).
In order to simplify our analysis processes, we make the following scale transformation:
We still denote by . Hence, system (8) reads
where . Note that the orbit direction of system (10) is reversed to that of system (8) because of the transformation (9) and . Furthermore, system (10) has a invariant region .
It is easy to see that system (10) is a slow-fast system with respect to fast variable .
Hence, rescaling , we get the slow system of (10) as
where is the slow variable. Similarly, let in system (10) and (11). We obtain the degenerate system
which is defined on the critical manifold
with , and the layer system
Clearly, the critical manifold consists of four normally hyperbolic submanifolds
Furthermore, it is clear that and are attracting; and are repelling; is a turning point, and is a generic fold point.
Through a straight calculation, we can get the following theorem about the existence and stability of equilibriums of system (10).
Theorem 1. For system (10), the following conclusions hold:
System(10)has three trivial equilibriums and . Furthermore, is a unstable node and and are saddle points if .
If , system(10)has a unique positive equilibrium satisfying , where . Furthermore is a stable node located in the right branch of function for and is an unstable node located in the left branch of function for ; see Fig. 1. Here
and
Note that the equilibriums in Theorem 1 are the spatially homogeneous equilibriums of model (3). In what follows, we will investigate the travelling waves about spatially homogeneous equilibrium and the periodic solutions of model (3). In other words, we will study the heteroclinic orbits about equilibrium , canard cycles and relaxation oscillation cycle of system (10).

3 Travelling waves analysis
In this section, we investigate the heteroclinic orbits of system (10) with , where is located in the right branch of function .
On , system (12) is reduced to
which allows assert that the value of decreases on segment and and increases on segment . Hence, system (14) has a stable equilibrium corresponding to and an unstable equilibrium corresponding to . Indeed, there exists a singular orbit from to .
With similar analysis, we can get the dynamics of (12) on . Hence, the dynamics of (10) and (11) with are shown in Fig. 2(a). Moreover, we have the following lemma about the heteroclinic orbits of positive equilibrium .
Lemma 1. For any fixed , there is such that for all , the following conclusions hold:
System(11)has a heteroclinic orbit from saddle to stable node and aheterocilinc orbit from unstable node to saddle .
System(11)has a heteroclinic orbit from saddle to stable node .
System(11)has infinitely many heteroclinic orbits from unstable node to stable node .
Proof. (i) Since is a normally hyperbolic manifold, then, based on the Finichel theory [12, 18], there is a slow manifold in the -neighborhood of when is sufficiently small. Furthermore, the equilibriums and lie in the slow manifold , and the flows on can be viewed as -order perturbation of the flows determined by (14). For (11) with , it is clear that the stable manifold of transversally intersects the unstable manifold of based on dimension counting. Indeed, this means that the singular orbit between and persists under perturbation .
Furthermore, we can similarly prove the heteroclinic orbit between unstable node and saddle point .
(ii) Clearly, there exists a singular orbit from to because of . Note that the intersection of and is transverse by dimension counting. According to the Finechel theory [12, 18], is perturbed to two dimensional stable manifold , and is perturbed to the unstable manifold of saddle . Furthermore, the intersection of and is persist because of the transverse intersection .
(iii) The proof process is similar to that of (ii).

Note that the picture of Lemma 1 is shown in Fig. 2(b). Furthermore, for system (8), we have the following remark.
Remark 1. Since the transformation (9), the flows of system (8) are reversed to that of system (10). Hence, the positive equilibrium of system (8) is unstable, and the heteroclinic orbits in Lemma 1 are also heteroclinic orbits of system (8) with oppositive direction.
Since is normally hyperbolic attracting manifold, then there exists a attracting slow manifold near for sufficiently small . Moreover, the heteroclinic orbits in Lemma 1 are transversal, then they persist on slow manifold Indeed, according to Remark 1, we have the following theorem for system (6).
Theorem 2. For any fixed , there exist and such that for all and , the following conclusions hold:
System(6)has a heteroclinic orbit from saddle to saddle and a heterocilinc orbit from saddle to stable node .
System(6)has a heteroclinic orbit from saddle to saddle .
System(6)has infinite heteroclinic orbits from saddle to stable node .
Note that these heteroclinic orbits in Theorem 2 are corresponded to different travelling waves of model (3). The biological explanation is that, if , the predators will be eventual extinction, and the prey will extinct or attach the carrying capacity. Furthermore, we give the following example to verify our conclusions.
Example 1. Set and in (8). It is clear that there exist the positive equilibrium and the heteroclinic orbits in Lemma 1; see Fig. 3. The travelling waves of model (3) are shown in Fig. 4, which correspond to the heteroclinic orbits in Theorem 2.


4 Periodic solutions analysis
In this section, we mainly investigate the existence of periodic solutions for model (3) with sufficiently small parameters . This means that we are interested in the periodic orbits of the slow-fast system (6). However, the canard explosion is the most important periodic orbit phenomenon in a slow-fast system. In [16, 17], this phenomenon is described as a small cycle arising in a singular Hopf bifurcation grows through a sequence canard cycles without head and canard cycles with head to a relaxation oscillation cycle. Canard cycles come from the perturbation of a singular slow-fast cycles, which consist of the attracting and repelling part of critical manifold and the fast orbits of layer system (13). Hence, we will give some results about the canard cycles, canard explosion phenomenon and relaxation oscillation cycle in this section.
To begin with, we give a result about the exit-entry function of system (10).
Lemma 2. For system (10)and fixed , there is an unique such that
Proof. Let
It is clear that and . Indeed, we also have . Hence, we can conclude that there is an unique such that .
4.1 Canard cycle and canard explosion
In order to find the canard cycles, we need to ensure that the generic fold point satisfies the canard point condition, i.e., , which means that the equilibrium of degenerate system (12) coincides with ; see Fig. 5(a). With a similar analysis, we obtain the dynamics of (10) and (11) with ; see Fig. 5(b).
Hence, we can construct a family of singular slow-fast cycles (see Fig. 6(a)) as
where , and are two roots of the equation , and a family of singular slow-fast cycles (see Fig. 6(b)) as


where defined by (15) in Lemma 2 when is the smaller root of the equation , and is the larger root of the equation .
For the dynamics of system (10) near canard point , we align to origin point by transformation and and denote by . Then we rewrite system (10) as
where
Note that we can choose suitable value of such that .
In order to use the conclusions in [16, 17], by a straightforward calculation, we can obtain the slow-fast normal form of (16) as
where
and
Clearly, implies that , which is equal to .
Hence, according to [16, 17], we obtain
and
where
Then the singular Hopf bifurcation curve and maximal canard curve of system (17) are
and
Since
then the equations and have unique solutions as
and
Hence, we have the following theorems about singular Hopf bifurcation from Theorem 3.1 in [17].
Theorem 3. For fixed , suppose the vertex point of curve is a canard point, and b and satisfy the relationship (18). Then there are and such that for each . and , system (16)has unique positive equilibrium in the small neighborhood of canard point and as . Furthermore, is stable with and unstable with , where is shown in(19). So a Hopf bifurcation occurs when b passes through It is supercritical if and subcritical if .
Proof. Let be the parameters satisfying . Hence, we have
Then the vertex point is a nondegenerate canard point. Furthermore, system (17) is the normal form of system (10) in the neighborhood of the nondegenerate canard point . Thus, it is clear that the origin point is a nondegenerate canard point of system (17).
According to Theorem 3.1 in [17], for suitable and , there is a equilibrium of system (17) near , which satisfy as . Furthermore, there exists a singular Hopf bifurcation curve such that is stable for and unstable for .
Since , then equation has unique solution , and is equal to . Hence, we complete the proof.
Moreover, from Lemma 2 and Theorems 3.3 and 3.5 in [17], the canard cycles are shown as follows.
Theorem 4. For fixed , suppose the vertex point of curve is a canard point, and b and satisfy th.e relationship(18). Then for and , there exists such that system (16)has a family of canard cycle emerges from ; see Fig. 6.
Moreover, the family uniformly converges to in Hausdorff distance as , and there exists a curve given in(20). For , the,canard explosion occurs when , where
Proof. The proof is omitted because its main idea is similar to that of Theorem 3.
Remark.2. Theorem 4 shows that, for and , there exists such that the singular slow-fast cycles or can be perturbed the canard cycles or of system (16). Furthermore, the canard explosion occurs when the√pa.ameter b changes in a exponential small neighborhood of the maximal canard curve .

Indeed, we give an example to show the canard explosion phenomenon as follows.
Example 2. Set in (10). It is clear that the value of maximal curve , and system (10) has the canard explosion phenomenon; see Fig. 7.
4.2 relaxation oscillation cycle
For , we have the dynamics of (10) and (11) with ; see Fig. 8(a).
Let be the intersection point of and , where defined by (15) in Lemma 2 when . We denote a singular slow-fast cycle as
which contains two slow segments and two fast segments; see Fig.8(b). Indeed, according to Theorem 3.2 in [3, 32], we have the following theorem about relaxation oscillation cycle.

Theorem 5. For any fixed , there exists such that for , system (10)has a unique relaxation oscillation cycle in the -neighborhood of , which is hyperbolic attracting; see Fig. 8(b). Furthermore, the cycle converges to in the Hausdorff distance as .
Remark 3. Due to transformation (9), the direction of canard cycles and relaxation oscillation cycle of system (8) is opposite to system (10).
Indeed, we give the following example to show the existence of relaxation oscillation cycle.
Example 3. Set and in (10). There exists a relaxation oscillation cycle; see Fig. 9.
Since the critical manifold is a normally hyperbolic attracting manifold, then for sufficiently small , there is a slow manifold near such that canard cycles and relaxation oscillation cycle persist on it. Hence, we have the following theorem about periodic solution of (3).
Theorem 6. There exist and such that for all and , the following conclusions hold:
For fixed , suppose the vertex point of curve is a canard point. Then there exists , such that model(3)has a family of periodic solutions corresponding to canard cycles.
For fixed , model(3)has a periodic solution corresponding to relaxation oscillation cycle.
Note that the picture of a periodic solution of model (3) is given in Fig. 10, which corresponds to the relaxation oscillation cycle shown in Fig. 9.


Remark 4. The periodic solution of system (3) in Fig. 9 means that the predators and preys will coexit since the population of preys sometimes is very low.
5 Conclusion
In this paper, we investigate the diffusive Leslie–Gower prey–predator model (3) under the assumption that both prey and predator diffuse small and prey grows much faster than predator. Thus, this prey–predator model is a singular problem with two small parameters. Using the travelling waves transformation, model (3) is transformed to a multiscale slow-fast system (6) with . Applying the geometric singular perturbation theory, we prove the existence of heteroclinic orbits of system (6) for , which correspond to the travelling waves of model (3). Indeed, we also prove the existence of relaxation oscillation cycle of system (10) for and the canard explosion phenomenon of system (10) when is a canard point. These periodic cycles imply the existence of periodic solutions of model (3). We also give some numerical examples to show illustration of our theoretical results. Our results have important theoretical significance for the biological control of pests and the conservation of biodiversity. Meanwhile, our method used in this paper can be applied into other similar models.
References
1. W. Abid, R. Yafia, M.A. Aziz-Alaoui, H. Bouhafa, A. Abichou, Global dynamics on a circular domain of a diffusion predator-prey model with modified Leslie–Gower and Beddington- DeAngelis functional type, Evol. Equ. Control Theory, .(2):115–129, 2015, https://doi.org/10.3934/eect.2015.4.115.
2. S. Ai, Y. Du, R. Peng, Traveling waves for a generalized Holling–Tanner predator–prey model, J. Differ. Equations, 263(11):7782–7814, 2017, https://doi.org/10.1016/j.jde.2017.08.021.
3. B. Ambrosio, M.A. Aziz-Alaoui, R. Yafia, Canard phenomenon in a slow-fast modified Leslie– Gower model, Math. Biosci., 295:48–54, 2018, https://doi.org/10.1016/j.mbs.2017.11.003.
4. A. Atabaigi, A. Barati, Relaxation oscillations and canard explosion in a predator–prey system of Holling and Leslie types, Nonlinear Anal., Real World Appl., 36:139–153, 2017, https://doi.org/10.1016/j.nonrwa.2017.01.006.
5. H. Cai, A. Ghazaryan, V. Manukian, Fisher-KPP dynamics in diffusive Rosenzweig– MacArthur and Holling–Tanner models, Math. Model. Nat. Phenom., 14(4):404–425, 2019, https://doi.org/10.1051/mmnp/2019017.
6. P.T. Cardin, M.A. Teixeira, Finichel theory for multiple time scale singular perutrbation problems, SIAM J. Appl. Dyn. Syst., 16(3):1425–1421, 2017, https://doi.org/10.1137/16M1067202.
7. J.B. Collings, The effect of the functional response on the bifurcation behavior of a mite predator–prey interaction model, J. Math. Biol., 36:149–168, 1997, https://doi.org/10.1007/s002850050095.
8. P. De Maesschalck, S. Schecter, The entry–exit function and geometric singular perturbation theory, J. Differ. Equations, 260(8):6697–6715, 2016, https://doi.org/10.1016/j.jde.2016.01.008.
9. Z.J. Du, J. Li, Geometric singular perturbation analysis to Camassa–Holm Kuramoto– Sivashinsky equation, J. Differ. Equations, 306:418–438, 2022, https://doi.org/10.1016/j.jde.2021.10.033.
10. Z.J. Du, J. Liu, Y.L. Ren, Traveling pulse solutions of a generalized Keller–Segel system with small cell diffusion via a geometric approach, J. Differ. Equations, 270:1019–1042, 2021, https://doi.org/10.1016/j.jde.2020.09.009.
11. A. Ducrot, Z.H. Liu, P. Magal, Large speed traveling waves for the Rosenzweig–MacArthur predator–prey model with spatial diffusion, Physica D, 415:132730, 14 pp, 2021, https://doi.org/10.1016/j.physd.2020.132730.
12. N. Fenichel, Geometric singular perturbation theory for ordinary differential equations, J. Differ. Equations, 31(1):53–98, 1979, https://doi.org/10.1016/0022-0396(79)90152-9.
13. A. Ghazaryan, V. Manukian, S. Schecter, Travelling waves in the Holling–Tanner model with weak diffusion, Proc. R. Soc. Lond., Ser. A, 471:20150045, 2015, https://doi.org/10.1098/rspa.2015.0045.
14. J.C. Huang, Y.J. Gong, S.G. Ruan, Bifurcation analysis in a predator-prey model with constant- yield predator harvesting, Discrete Contin. Dyn. Syst., Ser. B, 18(8):2101–2121, 2013, https://doi.org/10.3934/dcdsb.2013.18.2101.
15. J.C. Huang, S.H. Liu, S.G. Ruan, X.A. Zhang, Bogdanov-Takens bifurcation of codimension 3 in a predator-prey model with constant-yield predator harvesting, Commun. Pure Appl. Anal., 15(3):1041–1055, 2016, https://doi.org/10.3934/cpaa.2016.15.1041.
16. M. Krupa, P. Szmolyan, Extending geometric singular perturbation theory to nonhyperbolic points-fold and canard points in two dimensions, SIAM J. Math. Anal., 33(2):286–314, 2001, https://doi.org/10.1137/S0036141099360919.
17. M. Krupa, P. Szmolyan, Relaxation oscillation and canard explosion, J. Differ. Equations,174(2):312–368, 2001, https://doi.org/10.1006/jdeq.2000.3929.
18. C. Kuehn, Multiple Time Scale Dynamics, Appl. Math. Sci., Vol. 191, Springer, Cham, 2015, https://doi.org/10.1007/978-3-319-12316-5.
19. P.H. Leslie, Some further notes on the use of matrices in population mathematics, Biometrika, 35:213–245, 1948.
20. C.Z. Li, H.P. Zhu, Canard cycles for predator-prey systems with Holling types of functional response, J. Differ. Equations, 254(2):879–910, 2013, https://doi.org/10.1016/j.jde.2012.10.003.
21. W.S. Liu, Exchange lemmas for singular perturbation problems with certain turning points, J. Differ. Equations, 167(1):134–180, 2000, https://doi.org/10.1006/jdeq. 2000.3778.
22. W.S. Liu, Geometric singular perturbations for multiple turning points: Invariant manifolds and exchange lemmas, J. Dyn. Differ. Equ., 18:667–691, 2006, https://doi.org/10.1007/s10884-006-9020-7.
23. W.S. Liu, D.M. Xiao, Y. Yi, Relaxation oscillations in a class of predator–prey systems, J. Differ. Equations, 188(1):306–331, 2003, https://doi.org/10.1016/S0022-0396(02)00076-1.
24. V. Manukian, On traveling waves of Gray–Scott model, Dyn. Syst., 30(3):270–296, 2015, https://doi.org/10.1080/14689367.2015.1027177.
25. W.J. Ni, M.X. Wang, Dynamical properties of a Leslie-Gower prey-predator model with strong Allee effect in prey, Discrete Contin. Dyn. Syst., Ser. B, 22(9):3409–3420, 2017, https://doi.org/10.3934/dcdsb.2017172.
26. P.J. Pal, P.K. Mandal, Bifurcation analysis of a modified Leslie–Gower predator–prey model with Beddington–DeAngelis functional response and strong allee effect, Math. Comput. Simul., 97:123–146, 2014, https://doi.org/10.1016/j.matcom.2013.08.007.
27. H. Shen, C.H. Hsu, T.H. Yang, Fast–slow dynamics for intraguild predation models with evolutionary effects, J. Dyn. Differ. Equations, 32:895–920, 2020, https://doi.org/10.1007/s10884-019-09744-3.
28. J.H. Shen, Canard limit cycles and global dynamics in a singularly perturbed predator–prey system with non-monotonic functional response, Nonlinear Anal., Real World Appl., 31:146– 165, 2016, https://doi.org/10.1016/j.nonrwa.2016.01.013.
29. Y. Vera-Damin, C. Vidal, E. Gonzlez-Olivares, Dynamics and bifurcations of a modified Leslie–Gower-type model considering a Beddington–DeAngelis functional response, Math. Methods Appl. Sci., 42(9):3179–3210, 2019, https://doi.org/10.1002/mma.5577.
30. C. Wang, X. Zhang, Stability loss delay and smoothness of the return map in slow–fast systems, SIAM J. Appl. Dyn. Syst., 17(1):788–822, 2018, https://doi.org/10.1137/17M1130010.
31. C. Wang, X. Zhang, Canards, heteroclinic and homoclinic orbits for a slow-fast predator- prey model of generalized Holling type III, J. Differ. Equations, 267(6):3397–3441, 2019, https://doi.org/10.1016/j.jde.2019.04.008.
32. C. Wang, X. Zhang, Relaxation oscillations in a slow–fast modified Leslie–Gower model, Appl. Math. Lett., 87:147–153, 2019, https://doi.org/10.1016/j.aml.2018.07.029.
33. R. Yafia, M.A. Aziz-Alaoui, Existence of periodic travelling waves solutions in predator prey model with diffusion, Appl. Math. Model., 37(6):3635–3644, 2013, https://doi.org/10.1016/j.apm.2012.08.003.